• Aucun résultat trouvé

Coxeter groups, symmetries, and rooted representations

N/A
N/A
Protected

Academic year: 2021

Partager "Coxeter groups, symmetries, and rooted representations"

Copied!
11
0
0

Texte intégral

(1)

HAL Id: hal-01404050

https://hal.archives-ouvertes.fr/hal-01404050

Submitted on 28 Nov 2016

HAL

is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire

HAL, est

destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Olivier Geneste, Luis Paris

To cite this version:

Olivier Geneste, Luis Paris. Coxeter groups, symmetries, and rooted representations. Communications

in Algebra, Taylor & Francis, 2018, 46 (5), pp.1996-2002. �10.1080/00927872.2017.1363221�. �hal-

01404050�

(2)

Coxeter groups, symmetries, and rooted representations

O

LIVIER

G

ENESTE

L

UIS

P

ARIS

20F55

1 Introduction

Let Γ be a Coxeter graph, and let (W, S) be the Coxeter system of Γ . A symmetry of Γ is a permutation g of S such that m

g(s),g(t)

= m

s,t

for all s, t ∈ S , where (m

s,t

)

s,t∈S

is the Coxeter matrix of Γ . Let G be a group of symmetries of Γ. Then G is necessarily finite, and it can be viewed as a group of automorphisms of W . We denote by W

G

the subgroup of W fixed under the action of G. M¨uhlherr [7] and H´ee [4], independently of one another, proved that W

G

is a Coxeter group. None of them gave explicitly the Coxeter graph ˜ Γ which defines W

G

. However, a third proof, different from the other two, with an explicit description of ˜ Γ, is given in Crisp [2,

3].

Let Π = {

s

| s ∈ S} be a set in one-to-one correspondence with S , let V be the real vector space having Π as a basis, and let h., .i be the symmetric bilinear form on V defined by h

s

,

t

i = − cos(π/m

s,t

) if m

s,t

6= ∞ and h

s

,

t

i = −1 if m

s,t

= ∞. For every s ∈ S we define the linear transformation f

s

: V → V by f

s

(x) = x − 2 hx,

s

i

s

. Then the map S → GL(V ), s 7→ f

s

, induces a celebrated faithful linear representation f : W → GL(V), called the canonical representation of W (see Bourbaki [1]). In our context, the triple (V, h., .i, Π) will be called the canonical root basis of Γ .

Let G be a group of symmetries of Γ. Then G acts on V sending

s

to

g(s)

for all s ∈ S , and this action leaves invariant the canonical form h., .i. Hence, the canonical representation f : W → GL(V) is equivariant, in the sense that f (g(w)) = g◦ f (w) ◦ g

−1

for all g ∈ G and all w ∈ W , and therefore f induces a linear representation f

G

: W

G

→ GL(V

G

), where V

G

= {x ∈ V | g(x) = x for all g ∈ G}.

A naive question would be: Is f

G

: W

G

→ GL(V

G

) the canonical representation of

W

G

? A positive answer would provide a way to (re)prove that W

G

is a Coxeter group

and to determine the Coxeter graph of W

G

. Unfortunately, simple calculations show

that f

G

is not the canonical representation in general. Nevertheless, one can transpose

(3)

this question to a larger family of linear representations, the rooted representations introduced by Krammer [5,

6], and, in this context, the answer is yes. Our purpose is

to show that.

A root basis of Γ is a triple (V, h., .i, Π), where V is a finite dimensional real vector space, h., .i is a symmetric bilinear form on V , and Π = {

s

| s ∈ S} is a collection of vectors in V in one-to-one correspondence with S , that satisfies the following properties:

(a) h

s

,

s

i = 1 for all s ∈ S ; (b) for all s, t ∈ S, s 6= t, we have

h

s

,

t

i = − cos(π/m

s,t

) if m

s,t

6= ∞ , h

s

,

t

i ∈ (−∞, −1] if m

s,t

= ∞ ; (c) there exists χ ∈ V

such that χ(

s

) > 0 for all s ∈ S.

As mentioned above, if Π is a basis of V and h

s

,

t

i = −1 whenever m

s,t

= ∞ , then (V, h., .i, Π) is called the canonical root basis of Γ .

This definition is taken from Krammer’s thesis [5,

6]. It is both, a generalization of the

canonical spaces and canonical forms defined by Bourbaki [1], and a new point of view on the theory of reflection groups developed by Vinberg [8]. Note also that Condition (c) in the above definition often follows from Conditions (a) and (b), but not always (see Krammer [6, Proposition 6.1.2]).

Let (V, h., .i, Π) be a root basis of Γ . For every s ∈ S we define the linear transforma- tion f

s

: V → V by f

s

(x) = x − 2hx,

s

i

s

. The following theorem can be proved for any root basis in the same way as it is proved in Bourbaki [1] for the canonical root basis.

Theorem 1.1

(Krammer [5,

6])

The map S → GL(V) , s 7→ f

s

, induces a faithful linear representation f : W → GL(V) .

The representation f : W → GL(V) of Theorem

1.1

is called the rooted representation of W associated with (V, h., .i, Π).

Let G be a group of symmetries of Γ, and let (V, h., .i, Π) be a root basis. As for the

canonical root basis, we assume that G embeds in GL(V), satisfies g(

s

) =

g(s)

for

all g ∈ G and all s ∈ S, and leaves invariant the form h., .i . Then the representation

f : W → GL(V) is equivariant in the sense that f (g(w)) = g ◦ f (w) ◦ g

−1

for all g ∈ G

and all w ∈ W , and therefore f induces a linear representation f

G

: W

G

→ GL(V

G

),

where V

G

= {x ∈ V | g(x) = x for all g ∈ G}. The goal of this paper is to prove the

following.

(4)

Coxeter groups, symmetries, and rooted representations 3

Theorem 1.2

(1) The group W

G

is a Coxeter group.

(2) Let Γ ˜ denote the Coxeter graph of W

G

. There exists a subset Π ˜ of V

G

such that (V

G

, h., .i, Π) ˜ is a root basis of Γ, and the induced representation ˜ f

G

: W

G

→ GL(V

G

) is the rooted representation associated with (V

G

, h., .i, Π) ˜ . In particular, f

G

is faithful.

A similar approach is adopted in H´ee [4, Section 3], with different definitions. One can easily show that the root system obtained from a root basis is a root system in Hée sense [4], and that part of the results of the paper, such as the fact that (V

G

, h., .i, Π) is a ˜ root basis, can be deduced from Hée [4]. However, to get the explicit expression of the Coxeter graph ˜ Γ of W

G

, one would need extra arguments that can be either a rewrite of Lemma

3.3, or some arguments similar to that given in Crisp [2,3]. More generally,

the whole theorem is more or less in the literature. In particular, as mentioned before, Part (1) is explicit in M¨uhlherr [7], H´ee [4] and Crisp [2,

3]. But, our aim is to provide

a new point of view on the question with unified, short and self-contained proofs.

A more precise statement of Theorem

1.2

is given in Section

2. In particular, the

Coxeter graph ˜ Γ and the set ˜ Π are explicitly described. Section

3

is dedicated to the proofs.

2 Statement

The length of an element w ∈ W , denoted by lg(w), is the shortest length of an expression of w over the elements of S . An expression w = s

1

· · · s

`

is called reduced if ` = lg(w). It is known that, if W is finite, then W has a unique longest element, that is, an element w

0

∈ W such that lg(w) ≤ lg(w

0

) for all w ∈ W , and this element is an involution (see Bourbaki [1]).

For X ⊂ S, we denote by Γ

X

the full subgraph of Γ spanned by X , and by W

X

the subgroup of W generated by X. The subgroup W

X

is called a standard parabolic subgroup of W . By Bourbaki [1], (W

X

, X) is a Coxeter system of Γ

X

. If W

X

is finite, then we denote by w

X

the longest element of W

X

.

Let G be a group of symmetries of Γ. Now, we define a Coxeter matrix ˜ M = M ˜

G

= ( ˜ m

X,Y

)

X,Y∈S

(and its associated Coxeter graph, ˜ Γ). This will be the Coxeter matrix (and the Coxeter graph) of W

G

(see Theorem

1.2

and Theorem

2.2).

We denote by O the set of orbits of G in S , and we set S = {X ∈ O | W

X

is finite}.

Then S is the set of indices of ˜ M (which is the set of vertices of ˜ Γ ). Let X, Y ∈ S .

(5)

If m

s,t

= 2 for all s ∈ X and all t ∈ Y , then we set ˜ m

X,Y

= m ˜

Y,X

= 2.

Let k ∈ { 1, 2, 3, 4, 5 } . If Γ

X∪Y

is a disjoint union of copies of the Coxeter graph depicted in Figure

2.1

(k), where the vertices corresponding to x

1

, x

2

, . . . belong to X and the vertices corresponding to y

1

, y

2

, . . . belong to Y , then we set

˜

m

X,Y

= m ˜

Y,X

= m if k = 1, ˜ m

X,Y

= m ˜

Y,X

= 4 if k ∈ { 2, 3 } , ˜ m

X,Y

= m ˜

Y,X

= 8 if k = 4, and ˜ m

X,Y

= m ˜

Y,X

= 6 if k = 5. In this case we say that (X, Y ) is a bi-orbit of type k.

We set ˜ m

X,Y

= m ˜

Y,X

= ∞ in the remaining cases.

x1m y1 x1 y1 x2 x1 y1 y2 x2

(1) (2) (3)

x1 y1 4 y2 x2

x1

x2

x3

y1

(4) (5)

Figure 2.1: Bi-orbits.

The next lemma will be used in the definition of the set ˜ Π. It is well-known and can be easily proved using [1, Chapter V, Section 4, Subsection 8].

Lemma 2.1

Let (V, h., .i, Π) be a root basis of Γ . Suppose that W is finite and that Π spans V . Then h., .i is a scalar product, and (V, h., .i, Π) is the canonical root basis of Γ. In particular, Π is a basis of V .

We turn back to the hypothesis of Theorem

1.2, that is,

Γ is any Coxeter graph, G is a group of symmetries of Γ, and (V, h., .i, Π) is a root basis of Γ . We assume that G embeds in GL(V) so that the form h., .i is invariant under the action of G, and g(

s

) =

g(s)

for all s ∈ S and all g ∈ G.

Let X be an element of S , that is, an orbit of G in S such that W

X

is finite. Set Π

X

= {

s

| s ∈ X}, and denote by V

X

the linear subspace of V spanned by Π

X

, and by h., .i

X

the restriction of h., .i to V

X

× V

X

. By Lemma

2.1,

Π

X

is a basis of V

X

and h., .i

X

is a scalar product. Let a

X

=

P

s∈X

s

. Note that a

X

∈ V

G

, hence, by the above, a

X

6= 0 and ka

X

k > 0. We set ˜

X

=

kaaX

Xk

for all X ∈ S , and ˜ Π = Π ˜

G

= { ˜

X

| X ∈ S}.

The main result of the paper, with a precise statement, is the following.

Theorem 2.2

(1) The set S

W

= {w

X

| X ∈ S} generates W

G

, and (W

G

, S

W

) is a

Coxeter system of Γ ˜ .

(6)

Coxeter groups, symmetries, and rooted representations 5

(2) The triple (V

G

, h., .i, Π) ˜ is a root basis of Γ, and the induced representation ˜ f

G

: W

G

→ GL(V

G

) is the rooted representation associated with (V

G

, h., .i, Π). ˜ In particular, f

G

is faithful.

Remark

The proof of Part (1) of Theorem

2.2

uses the induced representation f

G

: W

G

→ GL(V

G

). Nevertheless, the conclusion of Part (1) is always true because there is always a root basis which satisfies the hypothesis of the theorem: the canonical root basis.

3 Proof

We assume given a Coxeter graph Γ, a root basis (V, h., .i, Π) of Γ , and a group G of symmetries of Γ . We assume that G embeds in GL(V), satisfies g(

s

) =

g(s)

for all g ∈ G and all s ∈ S, and leaves invariant the form h., .i .

Let f : W → GL(V) be the rooted representation of W associated with (V , h., .i, Π).

From now on, in order to simplify the notations, we will assume that W acts on V via f , and we will write w(x) in place of f (w)(x) for w ∈ W and x ∈ V . Lemmas

3.1

to

3.4

are preliminaries to the proof of Theorem

2.2. Lemma3.1

(1) is well-known. It is a direct consequence of M¨uhlherr [7, Lemma 2.8], and its proof can be found in the beginning of the proof of M¨uhlherr [7, Theorem 1.3]. Lemma

3.1

(2) is also know. Its proof is implicit in Crisp [2], but, as far as we know, it is not explicitly given anywhere else.

Lemma 3.1

(1) The group W

G

is generated by S

W

.

(2) We have (w

X

w

Y

)

m˜X,Y

= 1 for all X, Y ∈ S such that m ˜

X,Y

6= ∞ .

Proof

As mentioned above, the proof of Part (1) can be found in M¨uhlherr [7]. So, we only need to prove Part (2). Let X ⊂ S be such that Γ

X

is a disjoint union of vertices (i.e. Γ

X

has no edge). Then W

X

is finite and w

X

=

Q

s∈X

s. Let X = {s, t}

be a pair included in S such that m

s,t

= m < ∞ . Then W

X

is finite, w

X

= (st)

m2

if m is even, and w

X

= (st)

m−12

s if m is odd. Now, let X, Y ∈ S . If m

s,t

= 2 for all s ∈ X and t ∈ Y , then w

X

and w

Y

commute, hence (w

X

w

Y

)

2

= 1, as w

X

and w

Y

are both involutions. Suppose that (X, Y ) is a bi-orbit of type j, where j ∈ {1, 2, 3, 4, 5}.

Let Γ

1

, . . . , Γ

`

be the connected components of Γ

X∪Y

. For i ∈ {1, . . . , `}, we denote by Z

i

the set of vertices of Γ

i

, and we set X

i

= X ∩ Z

i

and Y

i

= Y ∩ Z

i

. We have w

X

=

Q`

i=1

w

Xi

and w

Y

=

Q`

i=1

w

Yi

. Moreover, using the above observation together

(7)

with Theorem

1.1, it is easily checked that (wXi

w

Yi

)

m˜X,Y

= 1 for all i. It follows that (w

X

w

Y

)

m˜X,Y

=

Q`

i=1

(w

Xi

w

Yi

)

m˜X,Y

= 1.

Lemma 3.2

Let X ∈ S . Then one of the following two alternatives holds.

(I) Γ

X

is a disjoint union of vertices (i.e. Γ

X

has no edge).

(II) There exists m ∈

N

, m ≥ 3, such that Γ

X

is a disjoint union of copies of the Coxeter graph depicted in Figure

2.1

(1).

Proof

For s ∈ X we set v

s

(X) = |{t ∈ X | m

s,t

≥ 3}|. Since W

X

is finite, the connected components of Γ

X

are trees (see Bourbaki [1]), hence there exists s ∈ X such that v

s

(X) ≤ 1. On the other hand, since G acts transitively on X , we have v

s

(X) = v

t

(X) for all s, t ∈ X . So, either v

s

(X) = 0 for all s ∈ X , or v

s

(X) = 1 for all s ∈ X . If v

s

(X) = 0 for all s ∈ X , then we are in Alternative (I). If v

s

(X) = 1 for all s ∈ X , then we are in Alternative (II).

Let X ∈ S . We say that X is of type I if Γ

X

satisfies Condition (I) of Lemma

3.2, and

that X is of type II

m

if Γ

X

satisfies Condition (II).

Lemma 3.3

Let X, Y ∈ S , X 6= Y . Then

h ˜

X

, ˜

Y

i = − cos(π/ m ˜

X,Y

) if m ˜

X,Y

6= ∞ , h˜

X

, ˜

Y

i ∈ (−∞, −1] if m ˜

X,Y

= ∞ .

Proof

Observe that, if m

s,t

= 2 for all s ∈ X and all t ∈ Y , then h ˜

X

, ˜

Y

i = 0 and

˜

m

X,Y

= 2. Hence, we can assume that there exist s ∈ X and t ∈ Y such that m

s,t

≥ 3.

Since G acts transitively on X and leaves invariant Y , it follows that, for all s ∈ X , there exists t ∈ Y such that m

s,t

≥ 3. Similarly, for all t ∈ Y , there exists s ∈ X such that m

s,t

≥ 3.

Recall that a

X

=

P

s∈X

s

, a

Y

=

P

t∈Y

t

˜

X

=

kaaX

Xk

, ˜

Y

=

kaaY

Yk

. Choose s ∈ X and set v

X

= |{t ∈ Y | m

s,t

≥ 3}| and p

X

=

P

t∈Y

h

s

,

t

i = h

s

, a

Y

i . Since G acts transitively on X and leaves invariant Y , these definitions do not depend on the choice of s . Similarly, choose t ∈ Y and set v

Y

= |{s ∈ X | m

s,t

≥ 3}| and p

Y

=

P

s∈X

h

t

,

s

i = h

t

, a

X

i. The hypothesis that there exist s ∈ X and t ∈ Y such that m

s,t

≥ 3 implies that v

X

≥ 1 and v

Y

≥ 1.

Let s ∈ X and t ∈ Y . If m

s,t

≥ 3, then h

s

,

t

i ≤ −

12

, and if m

s,t

= 2, then h

s

,

t

i = 0.

It follows that

(3–1) p

X

≤ − v

X

2 .

(8)

Coxeter groups, symmetries, and rooted representations 7

On the other hand, we have

(3–2) |X| v

X

= |Y| v

Y

.

This is the number of edges in Γ connecting an element of X with an element of Y . A direct calculation shows that

(3–3) ka

X

k =

p

|X| if X is of type I ,

p

|X|(1 − cos(π/m)) if X is of type II

m

. Finally, by definition of p

X

,

(3–4) ha

X

, a

Y

i = |X| p

X

.

Case 1: X and Y are of type I . Applying Equations (3–2), (3–3), and (3–4) we get

(3–5) h ˜

X

, ˜

Y

i = p

X

√ v

Y

√ v

X

.

Applying Equation (3–1) to this equality we get h ˜

X

, ˜

Y

i ≤ −

vXvY

2

. It follows that, if either v

X

≥ 4, or v

Y

≥ 4, or v

X

, v

Y

≥ 2, then h ˜

X

, ˜

Y

i ≤ −1. If v

X

= 1, v

Y

≥ 2 and p

X

≤ − cos(π/4) = −

1

2

, then, by Equation (3–5), h ˜

X

, ˜

Y

i ≤ − 1. If v

X

= 1, v

Y

= 3 and p

X

= − cos(π/3) = −

12

, then, by Equation (3–5), h ˜

X

, ˜

Y

i = −

3

2

= − cos(π/6).

In this case (Y, X) is a bi-orbit of type 5 and ˜ m

Y,X

= m ˜

X,Y

= 6. If v

X

= 1, v

Y

= 2 and p

X

= − cos(π/3) = −

12

, then, by Equation (3–5), h ˜

X

, ˜

Y

i = −

2

2

= − cos(π/4). In this case (Y , X) is a bi-orbit of type 2 and ˜ m

Y,X

= m ˜

X,Y

= 4. If v

X

= 1, v

Y

= 1 and p

X

= − cos(π/m) with m 6= ∞ , then, by Equation (3–5), h˜

X

, ˜

Y

i = − cos(π/m). In this case (Y, X) is a bi-orbit of type 1 and ˜ m

Y,X

= m ˜

X,Y

= m. Finally, if v

X

= v

Y

= 1 and p

X

≤ − 1, then, by Equation (3–5), h ˜

X

, ˜

Y

i = p

X

≤ − 1. In this case (Y, X) is a bi-orbit of type 1 and ˜ m

Y,X

= m ˜

X,Y

= ∞ .

Case 2: X is of type II

m

and Y is of type I . Applying Equations (3–2), (3–3), and (3–4) we get

(3–6) h ˜

X

, ˜

Y

i = p

X

√ v

Y

p

v

X

(1 − cos(π/m)) . Applying Equation (3–1) to this equality, we get

(3–7) h ˜

X

, ˜

Y

i ≤ −

√ v

X

v

Y

2

p

(1 − cos(π/m)) . If m ≥ 5, then

p

1 − cos(π/m) <

12

, hence, by Equation (3–7), h ˜

X

, ˜

Y

i ≤ − √ v

X

v

Y

−1. So, we can assume that m ∈ {3, 4}. Then we have

p

1 − cos(π/m) ≤

1

2

and, by Equation (3–7), h ˜

X

, ˜

Y

i ≤ −

vXvY

2

. It follows that, if either v

X

≥ 2, or v

Y

≥ 2, then

(9)

h ˜

X

, ˜

Y

i ≤ −1. If v

X

= 1, v

Y

= 1 and p

X

≤ − cos(π/4) = −

1

2

, then, by Equation (3–6), h ˜

X

, ˜

Y

i ≤ − 1. If v

X

= 1, v

Y

= 1, p

X

= − cos(π/3) = −

12

and m = 4, then, by Equation (3–6), h ˜

X

, ˜

Y

i = −

2+ 2

2

= − cos(π/8). In this case (Y , X) is a bi-orbit of type 4 and ˜ m

Y,X

= m ˜

X,Y

= 8. If v

X

= 1, v

Y

= 1, p

X

= − cos(π/3) = −

12

and m = 3, then, by Equation (3–6), h ˜

X

, ˜

Y

i = −

1

2

= − cos(π/4). In this case (Y , X) is a bi-orbit of type 3 and ˜ m

Y,X

= m ˜

X,Y

= 4.

Case 3: X is of type II

m

and Y is of type II

m0

. Applying Equations (3–2), (3–3) and (3–4) we get

h ˜

X

, ˜

Y

i = p

X

√ v

Y

p

v

X

(1 − cos(π/m))(1 − cos(π/m

0

)) . Applying Equation (3–1) to this equality we get

X

, ˜

Y

i ≤ −

√ v

X

v

Y

2

p

(1 − cos(π/m))(1 − cos(π/m

0

)) . Since

p

(1 − cos(π/m)) ≤

1

2

and

p

(1 − cos(π/m

0

)) ≤

1

2

, it follows that h ˜

X

, ˜

Y

i ≤

− √

v

X

v

Y

≤ −1.

Lemma 3.4

Let X ∈ S , and let x ∈ V

G

. Then w

X

(x) = x − 2hx, ˜

X

i ˜

X

.

Proof

Let Γ

0

be a Coxeter graph, and let (W

0

, S

0

) be its associated Coxeter system, such that W

0

is finite. Let w

00

be the longest element of W

0

, and let (V

0

, h., .i

0

, Π

0

) be the canonical root basis of Γ

0

. Then, by Bourbaki [1], w

00

0

) = −Π

0

.

Let X ∈ S . Recall that Π

X

= {

s

| s ∈ X}. By Lemma

2.1

and the above, we have w

X

X

) = −Π

X

, hence w

X

(a

X

) = −a

X

, therefore w

X

( ˜

X

) = −˜

X

.

Recall that V

X

denotes the linear subspace of V spanned by Π

X

. For all x ∈ V and all u ∈ W

X

there exists y ∈ V

X

such that u(x) = x + y. This is true by definition for all s ∈ X , hence it is true for all u ∈ W

X

. Let x ∈ V

G

. Let y ∈ V

X

be such that w

X

(x) = x + y . Let y =

P

s∈X

λ

s

s

be the expression of y in the basis Π

X

. For g ∈ G we have

x +

X

s∈X

λ

s

s

= w

X

(x) = g(w

X

)(g(x)) = g(w

X

(x))

= g(x) +

X

s∈X

λ

s

g(

s

) = x +

X

s∈X

λ

s

g(s)

,

hence λ

s

= λ

g−1(s)

for all s ∈ X . Since G acts transitively on X , it follows that λ

s

= λ

t

for all s, t ∈ X . So, there exists λ ∈

R

such that w

X

(x) = x + λa

X

= x + λ ka

X

X

.

(10)

Coxeter groups, symmetries, and rooted representations 9

We have

hx, ˜

X

i = hw

X

(x), w

X

( ˜

X

)i = hx + λ ka

X

k ˜

X

, −˜

X

i = −hx, ˜

X

i − λ ka

X

k , hence λ ka

X

k = −2hx, ˜

X

i . So, w

X

(x) = x − 2hx, ˜

X

i ˜

X

.

Proof of Theorem2.2

We have h˜

X

, ˜

X

i = 1 for all X ∈ S by definition. We have h ˜

X

, ˜

Y

i = − cos(π/ m ˜

X,Y

) if ˜ m

X,Y

6= ∞ ,

X

, ˜

Y

i ∈ (−∞, −1] if ˜ m

X,Y

= ∞ ,

by Lemma

3.3. Let

χ ∈ V

be such that χ(

s

) > 0 for all s ∈ S . Let ˜ χ : V

G

R

be the restriction of χ to V

G

. Then, for X ∈ S , ˜ χ( ˜

X

) =

ka1

Xk

P

s∈X

χ(

s

) > 0. So, (V

G

, h., .i, Π ˜

G

) is a root basis of ˜ Γ .

Let ( ˜ W, S) be a Coxeter system of ˜ ˜ Γ , where ˜ S = {˜ s

X

| X ∈ S} is a set in one- to-one correspondence with S . By Lemma

3.1, the map ˜

S → S

W

, ˜ s

X

7→ w

X

, induces a surjective homomorphism γ : ˜ W → W

G

. By Lemma

3.4, the composition

f

G

◦ γ : ˜ W → GL(V

G

) is the rooted representation associated with (V

G

, h., .i, Π). ˜ By Theorem

1.1, it follows that

f

G

◦ γ is injective, hence γ is an isomorphism. So, (W, S

W

) is a Coxeter system of ˜ Γ , and f

G

: W

G

→ GL(V

G

) is the rooted representation associated with (V

G

, h., .i, Π). ˜

References

[1] N Bourbaki, El´ements de math´ematique. Fasc. XXXIV. Groupes et alg`ebres de Lie.´ Chapitre IV: Groupes de Coxeter et syst`emes de Tits. Chapitre V: Groupes engendr´es par des r´eflexions. Chapitre VI: syst`emes de racines,Actualit´es Scientifiques et Indus- trielles, No. 1337, Hermann, Paris, 1968.

[2] J Crisp,Symmetrical subgroups of Artin groups,Adv. Math. 152 (2000), no. 1, 159–

177.

[3] J Crisp, Erratum to: “Symmetrical subgroups of Artin groups”, Adv. Math. 179 (2003), no. 2, 318–320.

[4] J-Y H´ee,Syst`eme de racines sur un anneau commutatif totalement ordonn´e,Geom.

Dedicata 37 (1991), no. 1, 65–102.

[5] D Krammer,The conjugacy problem for Coxeter groups,Ph. D. Thesis, Utrecht, 1994.

[6] D Krammer,The conjugacy problem for Coxeter groups,Groups Geom. Dyn. 3 (2009), no. 1, 71–171.

(11)

[7] B M ¨uhlherr, Coxeter groups in Coxeter groups,Finite geometry and combinatorics (Deinze, 1992), 277–287, London Math. Soc. Lecture Note Ser., 191, Cambridge Univ.

Press, Cambridge, 1993.

[8] E B Vinberg,` Discrete linear groups that are generated by reflections,Izv. Akad. Nauk SSSR Ser. Mat. 35 (1971), 1072–1112.

IMB, UMR 5584, CNRS, Univ. Bourgogne Franche-Comté, 21000 Dijon, France IMB, UMR 5584, CNRS, Univ. Bourgogne Franche-Comté, 21000 Dijon, France olivier.geneste@u-bourgogne.fr, lparis@u-bourgogne.fr

Références

Documents relatifs

However, in the case of an odd exponent the first condition is not sufficient: a group generated by reflections in the sides of a hyperbolic triangle with angles (π/2, π/5, π/5) has

Krammer [20], Digne [12] and independently Cohen–Wales [6] extended Krammer’s [20] constructions and proofs to the Artin groups associated with simply laced Coxeter graphs of

- A characterization of all representations of the canonical commutation relations which are quasi-equivalent to the Fock repre- sentation is given and a

varieties with trivial canonical divisor, klt singularities, K¨ahler-Einstein metrics, stability, holonomy groups, Bochner principle, irreducible holomorphic symplectic

They are maximal in the following sense: if H occurs as the vertex group of a finite splitting of G over virtually cyclic subgroups, in such a way that H admits a convex

Lemma 8.2. We define H to be the kernel of this map.. All the generators have a non-trivial action on the cell, except g 2,2. Its abelianization is isomorphic to Z/2 × Z/2.. Thus,

For N = 5, 6 and 7, using the classification of perfect quadratic forms, we compute the homology of the Voronoï cell complexes attached to the modular groups SL N ( Z ) and GL N ( Z

Linear algebraic groups and their homogeneous spaces have been thoroughly in- vestigated; in particular, the Chow ring of a connected linear algebraic group G over an